Sei sulla pagina 1di 12

A theoretical description of the polarization dependence of the sum frequency generation spectroscopy of the water/vapor interface

Angela Perry, Christine Neipert, Christina Ridley Kasprzyk, Tony Green, Brian Space et al. Citation: J. Chem. Phys. 123, 144705 (2005); doi: 10.1063/1.2046630 View online: http://dx.doi.org/10.1063/1.2046630 View Table of Contents: http://jcp.aip.org/resource/1/JCPSA6/v123/i14 Published by the American Institute of Physics.

Additional information on J. Chem. Phys.


Journal Homepage: http://jcp.aip.org/ Journal Information: http://jcp.aip.org/about/about_the_journal Top downloads: http://jcp.aip.org/features/most_downloaded Information for Authors: http://jcp.aip.org/authors

Downloaded 20 Mar 2012 to 130.102.158.22. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/about/rights_and_permissions

THE JOURNAL OF CHEMICAL PHYSICS 123, 144705 2005

A theoretical description of the polarization dependence of the sum frequency generation spectroscopy of the water/vapor interface
Angela Perry, Christine Neipert, Christina Ridley Kasprzyk, Tony Green, and Brian Spacea
Department of Chemistry, University of South Florida, Tampa, Florida 33620-5250

Preston B. Moore
Department of Chemistry and Biochemistry, University of the Sciences in Philadelphia, Philadelphia, Pennsylvania 19104

Received 26 May 2005; accepted 5 August 2005; published online 11 October 2005 An improved time correlation function TCF description of sum frequency generation SFG spectroscopy was developed and applied to theoretically describing the spectroscopy of the ambient water/vapor interface. A more general TCF expression than was published previously is presented it is valid over the entire vibrational spectrum for both the real and imaginary parts of the signal. Computationally, earlier time correlation function approaches were limited to short correlation times that made signal processing challenging. Here, this limitation is overcome, and well-averaged spectra are presented for the three independent polarization conditions that are possible for electronically nonresonant SFG. The theoretical spectra compare quite favorably in shape and relative magnitude to extant experimental results in the O u H stretching region of water for all polarization geometries. The methodological improvements also allow the calculation of intermolecular SFG spectra. While the intermolecular spectrum of bulk water shows relatively little structure, the interfacial spectra for polarizations that are sensitive to dipole derivatives normal to the interfaceSSP and PPP show a well-dened intermolecular mode at 875 cm1 that is comparable in intensity to the rest of the intermolecular structure, and has an intensity that is approximately one-sixth of the magnitude of the intense free O u H stretching peak. Using instantaneous normal mode methods, the resonance is shown to be due to a wagging mode localized on a single water molecule, almost parallel to the interface, with two hydrogens displaced normal to the interface, and the oxygen anchored in the interface. We have also uncovered the origin of another intermolecular mode at 95 cm1 for the SSP and PPP spectra, and at 220 cm1 for the SPS spectra. These resonances are due to hindered translations perpendicular to the interface for the SSP and PPP spectra, and translations parallel to the interface for the SPS spectra. Further, by examining the real and imaginary parts of the SFG signal, several resonances are shown to be due to a single spectroscopic species while the donor O u H region is shown to consist of three distinct species consistent with an earlier experimental analysis. 2005 American Institute of Physics. DOI: 10.1063/1.2046630
I. INTRODUCTION

Liquid water interfaces are ubiquitous and important in chemistry and the environment. Thus, with the advent of interface specic nonlinear optical spectroscopies, such interfaces have been intensely studiedboth theoretically19 and experimentally.1035 Sum frequency generation SFG spectroscopy is a powerful experimental method for probing the structure and dynamics of interfaces. SFG spectroscopy is one of several experimental methods that measure a second-order polarization, and the more common electronically nonresonant experiment is considered here. SFG spectroscopy is dipole forbidden in isotropic mediasuch as liquids. Contributions from bulk-allowed quadrapolar effects have been demonstrated to be negligible in some cases,36,37 but can be included if necessary.38 Interfaces serve to break the isotropic symmetry, and produce a dipolar second-order
a

Author to whom correspondence should be addressed. Electronic mail: space@cas.usf.edu

signal. The SFG experiment employs both a visible and infrared laser eld overlapping in time and space at the interface, and can be performed in the time or frequency domain.14,39 In the absence of any vibrational resonance at the instantaneous infrared laser frequency, a structureless signal due to the static hyperpolarizability of the interface is obtained.3,11,17 When the infrared laser frequency corresponds to a vibration at the interface, a resonant lineshape is obtained with a characteristic shape that reects both the structural and dynamical environment at the interface.2,40,41 In this paper, classical molecular dynamics MD methods are used to model the dynamics of the water/vapor interface. Two complementary theoretical approachesquantum corrected time correlation function TCF and instantaneous normal mode INM methodsuse the congurations generated by MD as input to describe the SFG spectrum of the interface, and to ascertain the molecular origin of the SFG signal; both INM and TCF methods rely on a suitable spectroscopic dipole and polarizability model. This dual ap 2005 American Institute of Physics

0021-9606/2005/123 14 /144705/11/$22.50

123, 144705-1

Downloaded 20 Mar 2012 to 130.102.158.22. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/about/rights_and_permissions

144705-2

Perry et al.

J. Chem. Phys. 123, 144705 2005

proach was demonstrated to be highly useful in understanding condensed phase spectroscopy of water, other liquids, and interfacesclassical mechanics, especially in the context of quantum-corrected TCFs, has proven to be surprisingly effective in modeling intramolecular vibrational spectroscopy.1,2 In particular, TCF methods have provided a quantitative description of the O u H stretching lineshape in ambient liquid water, and INM methods have served to identify the molecular motions that result in the observed signal; these complementary techniques are equally effective for modeling water interfacial spectroscopy.14,4248 An INM approximation to SFG spectroscopy is quantum mechanical by construction, but offers a limited dynamical description. As a result, in bulk water and other liquid state intramolecular lineshapes , INM intramolecular resonances are broader than their TCF counterparts, but have the same central frequency and integrated intensity. This observation suggests the intramolecular INM spectra represent an underlying spectral density that is dynamically motionally narrowed in the actual lineshape.40 This is also found to be the case here for SFG spectra in all polarization conditions. This result contrasts with a previous report by us,1 and evidence from the literature.3,4 Previous TCF and INM calculations of the SSP polarization SFG O u H stretching spectrum of the water/vapor interface were very noisy, and suggested the spectra had equal breadththus, suggesting motional narrowing effects were not apparent in the spectra. The success of an approximate, nondynamical, frequency domain technique,4 and the similarity of the spectra to those obtained using TCF methods,1,3 appeared to be further evidence of spectra that could be described in the inhomogeneously broadened limit.49 That method,4 however, contains an empirically adjustable line width that effectively accounts for some motional narrowing making it difcult to draw conclusions. Because of methodological advances, it is now possible to calculate well-averaged TCF and INM spectra, and they unambiguously demonstrate SFG O u H stretching lineshapes at least at the water/vapor interface are signicantly motionally narrowed to a degree reminiscent of the bulk.50,51 This observation suggests dynamical motional narrowing effects are important at interfaces, and the dynamics are best described as intermediate between the fast and slow modulation limits of motional narrowing. In the slow modulation inhomogeneously broadened limit, all frequency uctuations of the oscillator are represented in the lineshape.40,49,51 A recent study in the Shen group also suggested motional averaging effects may well be signicant in the SPS geometry, and, in that case, the free O u H stretching peak is greatly diminished. Although that study did not address motional narrowing, the presence of motional averaging suggests motional narrowing is important because it is due to fast reorientational motions within the vibrational relaxation time for the mode that would also be expected to result in motional narrowing. In order to obtain better TCF results, long time cross correlations between the system dipole and polarizability need to be followed. Because molecular simulations of interfaces in Cartesian space necessarily produce two interfaces, simulation times were limited to the molecular diffusion time

between interfaces so molecules could not contribute to the signal at both interfaces during one MD run.1 This leads to TCFs without long time decays that are difcult to Fourier transform accurately. In this work, a weak restraining potential is added that connes the molecules over time to the half of the simulation box they start in in the dimension normal to the interface without signicantly perturbing the relevant short time dynamics and average structure of the liquid that contributes to the interfacial spectroscopy; even though the molecular diffusion constant normal to the interface is changed, the molecule is only contributing to the spectrum while resident at the interface, and is free of any signicant external potential. This modication permits the calculation of TCFs out to arbitrarily long timesresulting in sharp spectra that include intermolecular spectral lineshapes. Surprisingly, a welldened intermolecular mode was found to be prominent in the spectrum.2 It is centered at 875 cm1, and is comparable in integrated intensity to the rest of the intermolecular lineshapethe lineshape also has an intensity that is approximately one-sixth of the magnitude of the intense free O u H stretching peak for spectra taken in polarization geometries that are sensitive to dipole derivatives normal to the interface SSP and PPP . Using instantaneous normal mode methods, the resonance is shown to be due to a wagging mode localized on individual water molecules. Water molecules contributing to this resonance are at a slight angle to the interface with their oxygen atoms anchored in the interface, and the hydrogen atoms wagging nearly normal to the interface. The presence of another population, aside from the free O u H stretch, of interfacial molecules was recently proposed via indirect evidence,6,34,35 and that hypothesis is strongly supported by this work. Here, we have directly observed a spectroscopically distinct species, and clearly identied the vibrational mode responsible for the lineshape. Thus, experimental setups that permit taking spectra at relatively long wavelengths could probe this mode as a complement to the information contained in the free and donor O u H stretching modes.5257 At lower frequencies, welldened hindered translational modes are found both parallel and perpendicular to the interface. The perpendicular modes are prominent in the polarization conditions sensitive to dipolar changes normal to the interface SPP and PPP while the parallel modes are more pronounced in the SPS geometry which is sensitive to motions along the interface. Further, some of the time domain expressions given earlier for SFG spectroscopy were only correct for the modulus of the SFG signal but not for the real and imaginary portions ,1,5 or were not entirely general and only valid at high intramolecular frequencies .3,38 Note, some of these works also equated the complex quantum TCF with the real classical TCF, further complicating matters.3,38 The correct general and exact expressions are given below and include both the resonant and nonresonant contributions. The previous expressions all give acceptable lineshapes at high frequencies for the modulus of the second-order signal. They are not correct for calculating the amplitude of the signal that can be detected in a heterodyne experiment,58 or by taking advantage of phase interference effects.10 Here, it

Downloaded 20 Mar 2012 to 130.102.158.22. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/about/rights_and_permissions

144705-3

SFG spectroscopy of the water/vapor interface

J. Chem. Phys. 123, 144705 2005

is shown by carefully examining the real and imaginary parts of the SFG signal, individual mode contributions to an observed lineshape can be identied. Using this approach, the free O u H and the newly discovered modes were identied as individual spectroscopic species one type of oscillator at the interface , and the donor O u H region consists of three distinct species. This last conclusion agrees with results from a careful deconvolution of O u H stretching signal in an earlier experimental work that also found three species each with approximately the same central frequency.11,59 Thus, as a prelude to more complex interfaces, this joint TCF/INM approach is applied to the water/vapor interface producing good agreement with the shape and relative amplitudes of SFG measurements for all independent polarization conditions.41 The theoretical expression in terms of a TCF for the SFG signal is also presentedincluding corrections from expressions published previously1,3,5 in Sec. II and Appendix A. The MD, dipole, many body polarization methods, and associated parameters are also summarized in Sec. II. The theoretical results, and their comparison to experiment, are discussed in Sec. III. The paper is concluded in Sec. IV.
II. MODELS AND METHODS

its polarizability tensorwhere the subscripts represent the vector and tensor components of interest, respectively. The operator evaluated at time t is the Heisenberg representation of the operator jk t = eiHt/ jkeiHt/ ; Tr represents the trace of the operators. It is convenient to proceed by rewriting the FourierLaplace transform in Eq. 2.1 as the Fourier transform of a correlation function that can then be interpreted in the classical limit, and quantum corrected. Evaluating the commutator in Eq. 2.1 gives Tr ,
i jk

t = C t C* t = 2iCI t , 2.2 t = CR t + i CI t .

Ct =

i jk

The SFG signal consists of nonresonant due to the static hyperpolarizability and resonant contributions that are important when the infrared laser frequency distribution is resonant with a vibrational transition at the interface. The signal intensity, is proportional to the square of their sum: 2 2 Res 2 + N Res and directly measures I SFG the modulus squared in the typical homodyne detected effectively monochromatic frequency domain experiment. The superscripts denote the resonant and nonresonant contributions, 2 is the susceptibility tensor. The respectively, and SFG proportionality constants include a factor of 2 and Fresnel factors.60 The vibrational information is contained in the resonant signal, and the nonresonant was found to be a negative constant in the O u H stretching region;3,11 this can still lead to the nonresonant contribution changing the frequency-dependent intensities through cross terms in the squared modulus signal. Through isotopic substitution, the nonresonant contribution can be measured independently, and this permits the deconvolution of the full signal to extract Res ; this deconvolution was done for the SSP polarization geometry at the water/vapor interface.11 The second-order response is given theoretically by a combination of resonant and nonresonant terms.3,16,61,62 The resonant terms can be grouped to give a simple expression in terms of the systems polarizability and dipole. A derivation including the resonant and nonresonant terms from perturbais tion theory is given in the Appendix.16,62 Thus Res given by5,61
Res

In Eq. 2.2 , the superscript asterisk is the complex conjugate, and the subscripts denote the real and imaginary parts of C t both of which are themselves real functions. The angle brackets are the trace of the operators divided by the partition function in the standard notation.63 In the classical limit, CR t becomes the classical cross-correlation function of the system dipole and polarizability tensor elements, i.e., lim 0CR t = CCl t = i jk t . Since only classical TCFs can be calculated using classical MD and TCF theory, the goal is to write the response function entirely in terms of the quantum-corrected classical TCF, CCl t . To proceed, the imaginary part of the one time correlation function is related in frequency space exactly to the real / 2 CR ; where the subscripts depart: CI = tanh note the Fourier transform of the real and imaginary parts of the complex function C t which is a real function of frequency, i.e., C = 1 2 dtei
t

CR t + iCI t = CR

+ CI

Using the result obtained in Eq. 2.2 for the trace in Eq. 2.1 gives
Res

2
0

dt ei tCI t , 2.3

CI t =

ei

tanh

/2 CR

i
0

dt ei tTr

jk

t .

2.1

Equation 2.3 demonstrates the SFG experiment probes the imaginary part of C t . Note, CI t is written in a form that can be calculated using the real part of the correlation functionwhich is obtainable from classical MD. Due to causality, the FourierLaplace transform gives a real and imaginary part in Eq. 2.3 as the cosine and sine transform of CI t , respectively. Equation 2.3 can be simplied by changing the order of integrationperforming the frequency domain integral rst. Dening the real and imaginary parts Res Res +i I : of Res = R = 2 tanh /2 CR = 2
0

In Eq. 2.1 , = e H / Q for a system with Hamiltonian H and partition function Q at reciprocal temperature = 1 / kT, and k is Boltzmanns constant. is the system dipole and

sin

t CI t dt,

2.4

Downloaded 20 Mar 2012 to 130.102.158.22. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/about/rights_and_permissions

144705-4

Perry et al.

J. Chem. Phys. 123, 144705 2005

2 = P

tanh

/2 CR +

2
0

cos

t CI t dt.

2.5

To obtain Eqs. 2.4 and 2.5 , the identity 0 ei t dt = iP 1 / + was used. P designates the principle part.64 Due to technical constraints in producing intense tunable infrared laser light, the focus of SFG experiments is curkT , and clasrently on high frequency spectra where sical mechanics is clearly invalid. Building on our previous work, the classical correlation function result, that is amenable to calculation using MD and TCF methods, is quantum corrected using a harmonic correction factor: CR / 2 coth / 2 .65,66 This correction factor = CCl is exact in relating the real part of the classical harmonic coordinate correlation function to its quantum mechanical counterpart. Here, we are using it to correct functions, the dipole and polarizability, that contain higher orders of the coordinates, and exact corrections for harmonic systems of this type are still possible, but not neededthe linear dipole and Placzek approximation are adequate.65 Using this result, the TCF approximation to the resonant part of the SFG spectrum, Res, takes the form
I

eliminated the real frequency contribution, and doubled the imaginary frequency part of the susceptibility. The other work3 used frequency domain perturbation theory,16,62 and divided the terms into resonant and nonresonant contributions then recast the resonant contribution as a TCF. This led to a TCF expression that replaced CI t with the full TCF, C t , in Eq. 2.3 to within a constant factor . Note, at high frequency, the tanh factor is almost unity, and their expression is a correct limiting expressionby making the rotating wave approximation the expression is only correct at high frequencies.61,69 In that work, if two of their nonresonant terms are included in the resonant contribution, then the exact Eq. 2.3 is obtainedthis result is demonstrated in the Appendix. At lower frequencies, the two expressions are quite different, and the tanh factor produces a time derivative of the correlation function in the time domain. To construct an INM approximation to Eq. 2.1 , it is sufcient to evaluate the trace in Eq. 2.1 for a harmonic system. It is convenient to invoke both the Placzek and linear dipole approximation to evaluate the resulting matrix elementsalthough higher-order contributions can be included, and simple analytic expressions result for these contributions. An equivalent approach is to evaluate CCl t for classical harmonic oscillators, and quantum correct the resulting expression using the harmonic correction factor, given above, to relate CCl t and CR t : CCl =
i/

= = P

CCl CCl

, d ,

2.6 2.7

Ql

jk/

Ql

kT
l 2

2.9

+
jk

CCl t =

t .

2.8

In Eq. 2.8 , the angle brackets represent a classical TCF that can be computed using MD, and a suitable spectroscopic model.67 Finally, Eqs. 2.6 and 2.7 give the TCF signal in a form amenable to classical simulation. Note, while it easier using Eq. 2.6 , R is more easily comto evaluate I puted by doing the cosine integral as in Eq. 2.5 . Considering the three possible independent polarization conditions, SSP, PPP, and SPS, for the TCF in Eq. 2.8 , the rst index in the polarization designation corresponds to the last index in the TCF. For example, the SSP and PPP polarization conditions probe dipolar motions normal to the interface, and the SPS case is sensitive to dipolar changes parallel to the interface. Note, the PPP condition is sensitive to motions both parallel and perpendicular to the interface.68 Further, the SSP and PPP / SPS probe diagonal/off-diagonal polarizability matrix elements, respectively. A similar TCF approach was adopted earlier by others3,5 and us1 for modeling the SFG spectrum of both solid5 and liquid interfaces1,3note, quantum corrections were not included in the works by the other groups. The earlier papers, and our previous work, did not give the exact expression for the real and imaginary parts of the SFG signal. Two of the papers,1,5 started with the time domain expression for the second-order response function,61 one improperly evaluated a contour integral which violated causality. This effectively

In Eq. 2.9 , l is the frequency of mode Ql, and the angle brackets represent averaging over classical congurations of is then back transformed into the system generated. CCl the time domain, and used in Eqs. 2.6 and 2.7 in place of the classical TCF to obtain an INM approximation to the spectroscopy. Below, it will be demonstrated the TCF approach, which does not invoke the Placzek and linear dipole approximation except implicitly in quantum correcting the results , gives results in close agreement with the INM results and Eq. 2.9 is therefore sufcient. MD simulations were performed using a code developed at the Center for Molecular Modeling at the University of Pennsylvania, and uses reversible integration and extended system techniques.70 Microcanonical MD simulations were performed on ambient H2O with a density of 1.0 g / cm3, and an average temperature of 298 K. To create an interface, a cubic simulation box of equilibrated liquid water was extended doubled along the z axis, and the system was allowed to equilibrate creating two water/vapor interfaces. The interfaces were sufciently far apart so as they did not interact strongly, and Ewald summation was included in three dimensions.9 The density prole of the system was monitored to verify equilibration.9 In all cases, the results were tested, and found to be system-size independent. Most results were generated from 216 molecule simulations, and smaller system sizes down to 64 molecules were tried, and did not alter the results.2 As in our previous work, MD simulations were conducted using a exible simple point charge SPC model that included a harmonic bending potential, linear cross terms

Downloaded 20 Mar 2012 to 130.102.158.22. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/about/rights_and_permissions

144705-5

SFG spectroscopy of the water/vapor interface

J. Chem. Phys. 123, 144705 2005

FIG. 1. Color online SFG SSP TCF spectra for the water/vapor interface highlighting the spectral changes in the use of two different Morse potentialsthe original Morse potential dashed blue line , and a softer Morse potential solid green line . The softer potential results in a shift of approximately 100 cm1 in the O u H stretching spectrum.

and Morse O u H stretching potentials, V r = De 1 e r 2.50 The Morse O u H stretching potential used here was slightly softer than previous work; our value is 2.50185 1 1 1,50 instead of 2.566 . For a Morse potential, the force constant, k, can be approximated as k = 2De 2. Assuming a harmonic oscillator with frequency = k / m, this implies the ratio 1 : 2 is proportional to the ratio 1 : 2. Therefore, a 2.5% change in the exponential Morse parameter implies a 2.5% shift in the spectral frequencies, and this behavior is demonstrated in Fig. 1. This analysis assumes the relevant coordinates are simple one-dimensional O u H stretching modes. If several distinctly different types of modes were present, a change in the shape of the broad O u H stretching signal would be expected. This is additional evidence that interfacial normal modes are well approximated as simple O u H stretches.1,2,4 Figure 1 highlights the spectral changes resulting from using a softer Morse potential. The slightly softer potential does not alter the intermolecular region of the spectra as would be expected. The intermolecular portion of the spectrum has polarizability and dipole derivatives changes that are due primarily to reorientation. These changes then depend on the polarizability tensor and the dipoles themselves, and not their derivatives. On the other hand, the intramolecular region of the spectra is simply shifted to the redthis point will be returned to when discussing the modal composition of the broad O u H stretching lineshape. This change resulted in the free O u H stretching frequency in better agreement with experimental values even though the Morse potential change is almost imperceptible to the naked eye. This implies the center of the lineshape is very sensitive to the local frequency along the Morse potential as the O u H stretching motion, perturbed by hydrogen bonding in the liquid, explores the highly anharmonic potential surface. In performing the MD, partial point charges were placed on the atoms that were chosen to reproduce the condensed phase dipole moment. At the water/vapor interface, the true water dipole falls from its condensed phase value, about 2.9

Debeye, to that in the gas phase, 1.8 Debeye, over a distance of only a few molecular layers.71 It would seem polarizable dynamics would be essential to model the dynamics of aqueous interfaces, but the use of nonpolarizable MD seems to adequately represent the structure of the water/vapor interface. A previous work using a polarizable model in this context is consistent with this observation.3 Evaluating the TCF in Eqs. 2.6 2.8 presents a problem for interfacial systems. The interface was constructed using the standard MD geometry with vacuum/vapor above and below the water.4,8 Unfortunately, this produces two interfaces with average net dipoles in opposite directions. Calculating the SFG spectrum of the entire system would lead to partial cancellation of the SFG signal, and meaningless results. Another problem arises in that molecules at one interface can diffuse to the other interface over time. In this case, simulation times are limited to the molecular diffusion time between interfaces so that molecules cannot contribute to signal at both interfaces during one MD run. This leads to TCFs without long time decays that are difcult to Fourier transform accurately.1,3 In order to obtain better TCF results, long time cross correlations between the system dipole and polarizability need to be followed. A weak laterally isotropic restraining potential was added effectively conning the molecules over time to the half of the simulation box they start in in the dimension normal to the interface without signicantly perturbing the relevant short time dynamics; even though the molecular diffusion constant normal to the interface is changed, the molecule is only contributing to the spectrum while resident at the interface, and is free of any signicant external potential. This modication permits the calculation of TCFs out to arbitrarily long times resulting in sharp spectra that include intermolecular spectral lineshapes. The restraining potential is of the form V = / r 9 with = 2.3 K, = 2.474 , and r = 0 is at the center of the box. The restraining potential becomes negligible near the interface, and is only signicant within 2.0 of the box center. The interfacial density prole was unchangeddemonstrating the restraining potential used did not perturb the average structure of the liquid that contributes to the interfacial spectroscopy. The MD was performed without explicit polarization forces; when the SFG TCF or INM spectrum are calculated, polarizability is included in the calculationsover 3 million 3 fs time steps were included in calculating the MD and TCFs. The model employed includes full many-body polarization effects included explicitly via a point atomic polarizability approximation PAPA polarizability model7274 with 3 point polarizabilities on the atoms O = 1.1482 , H 3 75 = 0.3304 . The permanent dipoles were calculated based on ab initio data as described previously.1,3 The SFG signal is sensitive to both dipole and polarizability derivatives. PAPA polarizability models naturally incorporate parameters that determine the polarizability derivatives. To implement this, it is sufcient to make the point polarizabilities on the atomic centers O u H bond-length dependent.7274 The r. r point polarizabilities then change as r = 0 r + is displacement from the equilibrium bond length. The

Downloaded 20 Mar 2012 to 130.102.158.22. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/about/rights_and_permissions

144705-6

Perry et al.

J. Chem. Phys. 123, 144705 2005

FIG. 3. Color online SFG SSP spectra for the water/vapor interface for the entire water vibrational spectrum using TCF solid green line method and INM dashed blue line method.

FIG. 2. Color online The a IR TCF spectra for liquid water, the b isotropic Raman TCF spectra for liquid water, and the c SFG SSP TCF spectra for the water/vapor interface highlighting the spectral changes in the use of two polarizability modelsprevious model dashed blue line and current model solid green line .
2 parameters for hydrogen and oxygen O = 2.7 , H= 2 1.06 are somewhat different than in our previous model, but still give reasonable values for the gas phase Raman and IR transition moments.1 Figure 2 c highlights the differences between the previous and current model for the SFG SSP TCF spectra showing the same SFG SSP spectrum presented in Fig. 1 . It is necessary, but not sufcient, to simply match the gas phase spectroscopic data, and suggests that interfacial molecules explore geometries different from both the gas phase and the bulk where the earlier polarizability model worked very well .45,47 Fitting to ab initio data for these interfacial geometries is clearly desirable, and is being pursued.3 Further, point atomic polarizability models, such as the one used here,1,45,72,75,76 offer exible and transferable parameters for both neat mixtures and liquids. They also offer a natural description of the induced dipole derivatives, and the ability to t polarizability derivatives.74 However, to produce an accurate description of interfacial polarizability derivatives, it may be necessary to make the point polarizabilities depend on bond angles, and not simply bond lengths as was done in this work.1 It is interesting to note the new model captures the free O u H mode more accurately without signicantly perturbing the intermolecular region of the spectraintramolecular spectra are sensitive to dipole and polarizability derivatives that do not signicantly change the magnitude of the dynamically more important dipole and polarizability. Note, the small differences in the intermolecular spectrum are likely due to the relatively poor averaging that was done in calculating the spectrum using the previous model. In this case only one-fteenth of the number of congurations were included in the calculation, and the SFG TCFs were slow to converge.1,38 Thus, even relatively small changes to these derivatives can greatly effect the spectroscopic observable

without changing the essential physics of the probleme.g., the identity of the relevant modes and motions. Figure 2 also presents the a infrared and b isotropic Raman TCF spectra relevant to the SSP polarization condition because it probes diagonal elements of the polarizability matrix for liquid water using both models. Again, only the intramolecular region of the spectra changed. For the O u H stretching region, increased asymmetry in the lineshape is apparent for the new model with a shoulder on the blue side. This is consistent with previous work that identied this shoulder to be due to instances in which a hydrogen does not form a hydrogen bond in the bulk;77,78 this would be analogous to the free O u H stretch found in interfacial spectra. The new polarizability model does a better job at highlighting this nonhydrogen bonded frequency distribution for liquid water, and, consequently, allows for more accurate interfacial spectra. The gure also clearly demonstrates the power of calculating spectroscopic observables to analyze condensed phase and interfacial structure. Interestingly, the shoulder on the blue side of the bulk Raman and IR spectrum is at the same central frequency as the free O u H mode at the interfacestrongly suggesting the presence of free, nonhydrogen bonded, O u H modes in bulk water.
III. DISCUSSION

Figure 3 displays the theoretical SFG SSP spectra for the entire water vibrational spectrum derived from both TCF and INM methods. Both the TCF and INM results are in absolute units, and no parameters were adjusted in displaying the data. The INM and TCF spectra were found to integrate to the same value over the entire 05000 cm1 range, and separately over the O u H stretching region 20005000 cm1 for all polarization conditions the others are not shown . This behavior is strong evidence for the interpretation of the INM lineshape as an underlying spectral density that is motionally narrowed in the observed spectrum.40 INM approximations to spectroscopy offer only a limited dynamical description, and correspond to an underlying spectral density that is typically broader than the observed lineshape when considering intramolecular modes. As an example, in bulk

Downloaded 20 Mar 2012 to 130.102.158.22. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/about/rights_and_permissions

144705-7

SFG spectroscopy of the water/vapor interface

J. Chem. Phys. 123, 144705 2005

FIG. 4. Color online SFG TCF spectra for the water/vapor interface in the O u H stretching region for three polarizations: SSP solid green line , PPP dashed blue line , and SPS dotted red line .

FIG. 5. Color online SFG TCF spectra for the water/vapor interface in the intermolecular region for three polarizations: SSP solid green line , PPP dashed blue line , and SPS dotted red line .

water and other liquid state intramolecular lineshapes INM intramolecular resonances were found to be broader than their TCF counterparts, but with the same central frequency and integrated intensity. Our TCF and INM spectra in Fig. 3 unambiguously demonstrate SFG O u H stretching lineshapes at the water/vapor interface are signicantly motionally narrowed to a degree reminiscent of the bulk.50,51 This result also suggests SFG spectra are sensitive to both structure and dynamics. The INM spectrum clearly exhibits the same resonances, but is broader. This implies the observed lineshapes are motionally narrowed, and dynamical contributions to SFG signals are important.41 Figure 4 presents TCF derived theoretical descriptions of the SFG spectra in the O u H stretching region for the water/ vapor interface. The three possible independent polarization conditions SSP, PPP, and SPS , in the electronically nonresonant experiment are displayed. The rst two indices can be interpreted as the element of the system polarizability tensor, and the second index as the element of the system dipole that is being probed. In the data, for all polarizations, we have included the SSP nonresonant contribution this is only strictly correct for the SSP polarization condition, and serves as an estimate in the other cases , N Res , which is a small negative constant,3,59 and the full signal is given by 2 2 Res 2 + N Res . In order to account for the SFG Fresnel coefcients that modify the experimental intensities, we have adjusted the relative intensities of our theoretical spectra so they can be more easily compared with experimental results.41 The spectrum in the SSP geometry that correlates the dipole moment component normal to the interface with diagonal polarizability matrix elements in the plane of the interface e.g., z 0 xx t , with the z axis taken as the surface normal direction leads to the most intense spectrum due to a relatively sizable, and changing, net normal dipole moment at the interface, and the relatively large diagonal polarizability elements; water has a nearly diagonal polarizability matrix with nearly equal elements in both the gas phase and bulk. Note, the PPP polarization condition is sensitive to a combination of all allowed susceptibility tensor elements in

contrast to SSP and SPS that only probe a single tensor element.68 Previously, we showed the agreement between the TCF and experimental spectrum, including the relative intensities of the different polarization conditions, was excellent, and within the statistical error over most of the frequency range.2 Thus, the essential features of the spectrum, and its polarization dependence, are captured very well by the TCF theory with the caveat that absolute intensities of the intramolecular modes are quite sensitive to the choice of polarizability parameters. The polarization dependence of the signal is demonstrated in Fig. 4. For polarizations that are sensitive to dipole derivatives normal to the interfaceSSP and PPPthe signal has an intense lineshape. In contrast, for the SPS geometry, which is sensitive to dipole derivatives parallel to the interface, only a hint of a signal is found. The SPS polarization condition also probes small off-diagonal polarizability matrix elements. These results also suggest by evaluation of the polarization dependence of the SFG spectra, given a knowledge of the expected nature of the polarizability and dipole derivatives, allows interfacial molecular geometries to be inferred via the spectra.10,68,79 While the intermolecular spectrum of bulk water shows little structure, the interfacial spectra are complex as shown in Fig. 5. The gure highlights the intermolecular SFG TCF spectra for the three independent polarization conditions SSP, PPP, and SPS . The polarizations that are sensitive to dipole derivatives normal to the interface, SSP and PPP, show a well-dened intermolecular mode at 875 cm1 that is comparable in intensity to the rest of the intermolecular structure and approximately one-sixth the intensity of the intense free O u H stretching peak.2 Using instantaneous normal mode methods looking at the nature of the INMs in the same spectral region , the resonance is shown to be due to a wagging mode localized on a single water molecule, at a slight angle to the interface, with two hydrogens vibrating/ librating normal to the interface, and the oxygen anchored in the interface.2 The hydrogens, pointing into the vapor phase, are hydrogen bonded to an oxygen atom at the interface. The SSP and PPP also show an intense intermolecular

Downloaded 20 Mar 2012 to 130.102.158.22. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/about/rights_and_permissions

144705-8

Perry et al.

J. Chem. Phys. 123, 144705 2005

FIG. 7. The probability distribution of the direction cosine from the surface normal of O u H vectors pointing into the vapor. FIG. 6. Color A snapshot of a water/vapor interface containing 216 water molecules featuring INMs from different regions of the spectra. The water molecule shown in blue is representative of a free O u H mode at 3694 cm1. The water molecule shown in green is representative of a wagging motion at 858 cm1. The water molecule shown in yellow highlights a translation perpendicular to the interface at 46 cm1. The water molecule shown in black highlights a translation parallel to the interface at 197 cm1.

mode at 95 cm1. Using instantaneous normal mode methods the resonance is found to be due to translations perpendicular to the interface. The SPS spectra, which are sensitive to dipole derivatives parallel to the interface, show an intermolecular mode at 220 cm1. This mode is a result of translations parallel to the interface. The importance of polarization sensitivity in SFG experiments is, thus, highlighted. Further, we have observed spectroscopically distinct species, and clearly identied the vibrational modes responsible for the lineshape. Hence, experimental setups that permit taking spectra at relatively long wavelengths could probe these modes as a compliment to the information contained in the free and donor O u H stretching modes.52,54 These three distinct populations of water molecules at the interface were previously undescribedother works have inferred the existence of something like the wagging mode.6,34,35 This might be considered surprising given the large numbers of MD simulations of the water/vapor interface that have been performed previously. This observation highlights the power of calculating spectroscopic observables in assessing interfacial structure and dynamics. Not only can the results be directly compared with experiment, thus validating the MD model, the spectroscopic calculation serves as a lter of the dynamics extracting out the identity of collective coordinates with well-dened frequencies that persist at the interface. Figure 6 highlights the vibrational modes from the intermolecular and intramolecular region of the spectra. A typical free O u H mode, shown in blue, produces the highfrequency feature at 3700 cm1. It is clear the oxygen atom is anchored in the interface, and the O u H is oscillating freely above the interface. The wagging mode giving rise to the spectral feature at 875 cm1 is displayed in green at the opposite interface. Here, the oxygen atom is anchored in the interface, and the two hydrogens are vibrating into the vapor phase. A representative perpendicular translational mode

with lineshape centered at 95 cm1 is shown in yellow, and the roughly parallel translational mode with lineshape centered at 220 cm1 is shown in black. These results demonstrate how INM approach does not require a priori assumptions about the nature of interfacial modes but does reveal their physical characteristics, and how different molecular motions contribute to the spectrum. In future work, a quantication of the relative populations of these interfacial species is planned via this approach. Figure 7 displays the distribution of the direction cosine from the surface normal of O u H vectors pointing into the vapor. This result compares well with previous theoretical 1. We data.4 We see an enhancement in probability at cos also nd approximately 20% of surface water molecules have a free O u H bond pointing out of the liquid, and into the vapor which is consistent with previous theoretical work.4,8 This analysis also points out it is necessary to talk of broad distributions of angles at the water/vapor interfaces, and that relatively less can be learned from single average values of orientations. Figure 8 a displays the real and imaginary parts for the SSP spectrum calculated via Eqs. 2.5 and 2.6 . Examining

FIG. 8. Color online Real solid green line and imaginary dashed blue line components of the a SFG SSP TCF spectra for the water/vapor interface and for b bulk water calculated as the FourierLaplace transform.

Downloaded 20 Mar 2012 to 130.102.158.22. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/about/rights_and_permissions

144705-9

SFG spectroscopy of the water/vapor interface

J. Chem. Phys. 123, 144705 2005

FIG. 9. Color online Real solid green line and imaginary dashed blue line components of the SFG SSP TCF spectra for the water/vapor interface for the O u H stretching region. The arrows highlight three separate modes centered at 3195 cm1, 3325 cm1, and 3400 cm1.

the real and imaginary parts of the spectrum can offer insights unavailable from the modulus alone. The real and imaginary parts could be measured experimentally via a heterodyne detection scheme, or by taking advantage of interference effects between bulk and interfacial contributions to the spectrum.10 To see the advantages of separately examining the real and imaginary contributions, it is useful to write the resonant SFG signal of a single harmonic mode, Q with linear dipole and polarizability , in frequency space as4
2 R i/

jk/


IR

IR 2

3.1

2 I

i/

jk/

IR

3.2

In Eqs. 3.1 and 3.2 , is a mathematical convergence parameter that physically can be interpreted as a homogeneous line width. The signal magnitude is seen to be proportional to the product of dipole and polarizability derivatives. Equations 3.1 and 3.2 imply a single type of mode will lead to an imaginary contribution that is a symmetric welldened peak Lorentzian in character while the real part will change sign, dipping below zero, at the maximum of the imaginary portion. If more than one species is contributing to the signal in a given region, a more complex lineshape will result from the overlapping signals. Examining the real and imaginary contributions in Fig. 8 a , it is clear several of the resonances are essentially single mode in character: the free O u H 3700 cm1 , the small bending contribution at the surface 1800 cm1 , the wagging mode 875 cm1 , and translational modes 95 cm1 and 220 cm1 . There is some overlap in the translational modes, and it is instructive the higher frequency 220 cm1 mode, that is pronounced only in the SPS modulus spectrum, also shows up in the SSP real and imaginary spectra. Figure 9 highlights the O u H stretching regionfrom approximately 3000 cm1 to 3600 cm1. Careful examination of the spectrum reveals three separate modes in this region centered at 3195 cm1, 3325 cm1, and 3400 cm1.

Remarkably, this agrees very well with previous experimental work that deconvoluted the spectrum in this region. That analysis revealed three modes present in the same region centered at 3200 cm1, 3325 cm1, and 3454 cm1nearly the same frequencies.11,59 This is strong evidence for distinct populations of water molecules in this donor O u H region of the spectrum. Further work is underway to identify the nature of these distinct O u H stretching species. It should be noted, that while the real and imaginary parts of the TCF derived SFG spectra do clearly indicate the presence of distinct subpopulations of oscillators, it is difcult to unambiguously identify the species responsible for the signals. This complication occurs because the modes are identied using INM methods, and the INM signal is broad in this region. Therefore, there is not an absolute correspondence between an INM frequency and the associated TCF spectrum in this congested spectral region . This difculty does not arise, however, in investigating spectra regions dominated by a single resonance like the free O u H or wagging mode. Further, the theoretical and experimental spectra have a somewhat different shape in this region, and this manifests itself in the relative intensities of the different contributions considering the extant water/vapor SFG spectra that have similar features but not identical shapes in this region .11,17,41,59,80 The differences are most likely due to the spectroscopic intensities of these species via our spectroscopic model rather than different populations of these species at the interface within the MD model. However, further investigation is required to denitively demonstrate this. It should also be noted, as pointed out in an earlier work,4 orientational information can also be deduced from the relative signs of the imaginary mode lineshapes given knowledge of the the signs of the prefactors in Eqs. 3.1 and 3.2 the dipole and polarizability derivatives . To further show the utility of the real and imaginary modal analysis, Fig. 8 b displays the real and imaginary parts of the bulk water O u H stretching region calculated as the FourierLaplace transform. While a linear IR experiment does not measure this observable, the transform can still be applied as an analysis tool. Figure 8 b is strong evidence there are two distinct species in the bulk, and the higher frequency moiety arises from the bulk free O u H, nonhydrogen bonded molecules.77,78,81
IV. CONCLUSION

The combined use of improved TCF and INM approximations to SFG spectroscopy represent a powerful complementary approach. Achieving agreement with experimental measurements engenders condence in the MD and spectroscopic models used to produce the theoretical spectrum. Many MD simulations of the water/vapor interface have been performed, but traditional analysis techniques do not easily uncover important interfacial subpopulations such as the wagging hindered rotational and hindered translational motions. Thus, SFG spectroscopy may be capable of giving a complete picture of the interfaceincluding structure and dynamics. Realizing this promise depends critically on the spectra being reliably interpreted, and the methods employed

Downloaded 20 Mar 2012 to 130.102.158.22. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/about/rights_and_permissions

144705-10

Perry et al.

J. Chem. Phys. 123, 144705 2005

in this study are designed to unambiguously characterize the nature of SFG spectra including inferring subpopulations of molecules from complex lineshapes. The plan is to investigate more complex and interesting interfaces using our improved combined INM/TCF approach.
ACKNOWLEDGMENTS

The resonant contributions can be simplied by rewriting in terms of polarizabilities and dipoles, and approximation 1 / SFG = 1 / vis. Given the denition of polarizability in Eq. A2 , the two resonant terms, R1 and R2, simplify to A3 and A4 , respectively:
pq

=
g,n

p gn

q ng ng

The research at USF was supported by an NSF Grant No. CHE-0312834 and a grant from the Petroleum Research Foundation to B.S., a Latino Graduate Fellowship to C.N., and a USF Presidential Fellowship to C.R.K. Acknowledgment is made to the Donors of the American Chemical Society Petroleum Research Fund for support of this research. The authors would like to acknowledge the use of the services provided by the Research Oriented Computing center at USF. The authors also thank the Space Foundation for Basic and Applied Research for partial support.
APPENDIX: DERIVATION OF A TIME DOMAIN EXPRESSION FOR THE SUM FREQUENCY RESPONSE

ng i

q gn

p ng ng

ng + i

0 g

, A2

R1 =

pq gm IR mg

r mg

+i

,
mg

A3

R2 =

r pq gn ng IR + ng

+i

.
ng

A4

Starting from frequency domain perturbation theory for , six terms are obtained, and are shown below.3,16,62 Four of the terms contribute to the resonant signal contained in R1 and R2 below , and two are nonresonant NR1 and NR2 below . While two of the resonant terms may appear initially to be nonresonant, those with denominators containing the expression IR + ng + i ng , they contribute to the resonant susceptibility, and lead to the complex conjugate correlation function C* t . Ultimately, their inclusion is necessary to reproduce Eq. 2.3 . In the expressions below, IR and vis are the frequencies of the infrared and visible elds. SFG is the sum frequency of the infrared and visible elds, and ng is the frequency corresponding to the energy difference between energy levels n and g. In Eq. A1 , g0 is the initial state thermal population, and the sum is over vibronic levels. and for , is a dipole matrix element between states dipole vector component :
2 SFG pqr SFG, vis, IR

Let Res denote only the sum of the resonant terms R1 pqr and R2. Replacing the denominators in both of the resonant terms with the integral identities 0 dt eit oi = i / o i and 0 dt eit + o+i = i / + o + i , and then taking the implicit limit that gamma goes to zero gives Eq. A5 . Equation A6 follows as an exact rewrite of Eq. A5 , and expresses the susceptibility in terms of the cross correlation of the system dipole and polarizability:
Res pqr =

i
gm 0

ei ei
ng 0

mgt

ei

IRt

pq gm

r mgdt

ngt

ei

IRt

pq ng

r gndt

0 g

A5

Res pqr =

i
0

dt ei i
0

IR

pq

dt ei

IR

pq

t .

A6

=
g,n,m

0 g

R1 + R2 + NR1 + NR2 ,
q gn vis + p nm ng

Expansion of the correlation functions in Eq. A6 results in a signicant simplication. Noting, r 0 pq t * gives Eq. A7 below = CR t + iCI t = pq t r 0 which is given in the text as Eq. 2.3 .
Res pqr IR

R1 =

p gn SFG

q nm ng

+i +i

ng

+i

2
0

dt e it

IR

CI t .

A7

ng

r mg IR q nm SFG + mg mg

,
p nm vis mg q mg

R2 =

p mg mg

+i +i

mg

+i

mg

r gn IR + ng ng p mg mg q mg

,
r nm vis + r nm vis mg ng

A1

NR1 =

q gn SFG + mg

+i
p gn

+i

,
ng

NR2 =

SFG

ng

+i

ng

+i

.
mg

A. Perry, H. Ahlborn, P. Moore, and B. Space, J. Chem. Phys. 118, 8411 2003 . A. Perry, C. Neipert, C. Ridley, and B. Space, Phys. Rev. E 71, 050601 1 2005 . 3 A. Morita and J. T. Hynes, J. Phys. Chem. B 106, 673 2002 . 4 A. Morita and J. T. Hynes, Chem. Phys. 258, 371 2000 . 5 V. Pouthier, P. Hoang, and C. Girardet, J. Chem. Phys. 110, 6963 1999 . 6 I. W. Kuo and C. J. Mundy, Science 303, 658 2004 . 7 P. Vassilev, C. Hartnig, M. T. Koper, F. Frechard, and R. A. van Santen, J. Chem. Phys. 115, 9815 2001 . 8 I. Benjamin, Phys. Rev. Lett. 73, 2083 1994 . 9 R. S. Taylor, L. X. Dang, and B. C. Garrett, J. Phys. Chem. 100, 11720 1996 . 10 V. Ostroverkhov, G. A. Waychunas, and Y. R. Shen, Phys. Rev. Lett. 94, 046102 2005 . 11 E. A. Raymond, T. L. Tarbuck, and G. L. Richmond, J. Phys. Chem. B 106, 2817 2002 .
2

Downloaded 20 Mar 2012 to 130.102.158.22. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/about/rights_and_permissions

144705-11
12 13

SFG spectroscopy of the water/vapor interface


47

J. Chem. Phys. 123, 144705 2005 H. Ahlborn, X. Ji, B. Space, and P. B. Moore, J. Chem. Phys. 111, 10622 1999 . 48 P. B. Moore, X. Ji, H. Ahlborn, and B. Space, Chem. Phys. Lett. 296, 259 1998 . 49 P. Moore and B. Space, J. Chem. Phys. 107, 5635 1997 . 50 H. Ahlborn, X. Ji, B. Space, and P. B. Moore, J. Chem. Phys. 111, 10622 1999 ; see discussion and references within. 51 C. P. Lawrence and J. L. Skinner, J. Chem. Phys. 117, 5827 2002 . 52 A. Tadjeddine and A. Le Rille, Electrochim. Acta 45, 601 1999 . 53 O. Pluchery and A. Tadjeddine, J. Electroanal. Chem. 500, 379 2001 . 54 D. K. Hore, M. Y. Hamamoto, and G. Richmond, J. Chem. Phys. 121, 12589 2004 . 55 D. Hore, J. King, F. Moore, D. Alavi, M. Hamamoto, and G. L. Richmond, Appl. Spectrosc. 58, 1377 2004 . 56 R. Braun, B. Casson, C. Bain, E. van der Ham, Q. Vrehen, E. Eliel, A. Briggs, and P. Davies, J. Chem. Phys. 110, 4634 1999 . 57 A. Bonvalet, J. Nagle, V. Berger, A. Migus, J. Martin, and M. Joffre, Phys. Rev. Lett. 76, 4392 1996 . 58 S. Mukamel, Principles of Nonlinear Optical Spectroscopy Oxford University Press, Oxford, 1995 , Chap. 5 pp. 129-130. 59 E. A. Raymond and G. L. Richmond, J. Phys. Chem. B 108, 5051 2004 . 60 Y. R. Shen, Annu. Rev. Psychol. 40, 327 1989 . 61 S. Mukamel, Principles of Nonlinear Optical Spectroscopy Oxford University Press, Oxford, 1995 . 62 R. W. Boyd, Nonlinear Optics Academic, London, 2003 . 63 D. A. McQuarrie, Statistical Mechanics Harper and Row, New York, 1976 . 64 W. Heitler, The Quantum Theory of Radiation Dover, New York, 1984 . 65 H. Kim and P. J. Rossky, J. Phys. Chem. B 106, 8240 2002 . 66 In our previous paper Ref. 1 we accidentally showed the harmonic quantum correction factor appropriate for the whole correlation function = CCl / 1 e instead of the correction for the real part, C CR , given here. Note that in the high frequency limit the two factors only differ by a factor of 2. 67 M. P. Allen and D. J. Tildesley, Computer Simulation of Liquids Clarendon, Oxford, 1989 . 68 X. Zhuang, P. B. Miranda, D. Kim, and Y. R. Shen, Phys. Rev. B 59, 12632 1999 . 69 J. S. R. J. Silbey, J. Chem. Phys. 115, 9266 2001 . 70 M. Tuckerman, B. J. Berne, and G. J. Martyna, J. Chem. Phys. 97, 1990 1992 . 71 T.-M. Chang and L. X. Dang, J. Chem. Phys. 104, 6772 1996 . 72 K. A. Bode and J. Applequist, J. Phys. Chem. 100, 17820 1996 . 73 J. Applequist, J. R. Carl, and K.-K. Fung, J. Am. Chem. Soc. 94, 2952 1972 . 74 J. Applequist and C. O. Quicksall, J. Chem. Phys. 66, 3455 1977 . 75 M. Souaille and J. C. Smith, Mol. Phys. 87, 1333 1996 . 76 B. Thole, Chem. Phys. 59, 341 1981 . 77 C. P. Lawrence and J. L. Skinner, J. Chem. Phys. 117, 8847 2002 . 78 C. P. Lawrence and J. L. Skinner, J. Chem. Phys. 118, 264 2003 . 79 J. Wang, Z. Paszti, M. A. Even, and Z. Chen, J. Am. Chem. Soc. 124, 7016 2002 . 80 C. Schnitzer, S. Baldelli, D. J. Campbell, and M. J. Shultz, J. Phys. Chem. A 103, 6383 1999 . 81 K. B. Moller, R. Rey, and J. T. Hynes, J. Phys. Chem. A 108, 1275 2004 .

Y. Shen, Solid State Commun. 108, 399 1998 . G. Richmond, Annu. Rev. Phys. Chem. 52, 357 2001 . 14 G. Richmond, Chem. Rev. Washington, D.C. 102, 2693 2002 . 15 D. E. Gragson and G. L. Richmond, J. Phys. Chem. B 102, 569 1998 . 16 Y. Shen, Principles of Nonlinear Optics Wiley, New York, 1984 . 17 Q. Du, R. Superne, E. Freysz, and Y. Shen, Phys. Rev. Lett. 70, 2313 1993 . 18 C. Raduge, V. Pumio, and Y. Shen, Chem. Phys. Lett. 274, 140 1997 . 19 Y. R. Shen, Proc. Natl. Acad. Sci. U.S.A. 93, 12104 1996 . 20 E. Freysz, Q. Du, and Y. R. Shen, Ann. Phys. Paris 19, 95 1994 . 21 Q. Du, E. Freysz, and Y. Shen, Science 264, 826 1994 . 22 R. Superne, J. Y. Huang, and Y. R. Shen, Phys. Rev. Lett. 66, 1066 1991 . 23 D. Gragson and G. Richmond, J. Chem. Phys. 107, 9687 1997 . 24 M. G. Brown, E. A. Raymond, H. C. Allen, L. F. Scatena, and G. L. Richmond, J. Phys. Chem. A 104, 10220 2000 . 25 D. E. Gragson and G. L. Richmond, J. Phys. Chem. B 102, 3847 1998 . 26 D. E. Gragson and G. L. Richmond, J. Am. Chem. Soc. 120, 366 1998 . 27 L. F. Scatena, M. G. Brown, and G. L. Richmond, Science 292, 908 2001 . 28 S. Baldelli, C. Schnitzer, M. J. Shultz, and D. J. Campbell, J. Phys. Chem. B 101, 10435 1997 . 29 S. Baldelli, C. Schnitzer, M. J. Shultz, and D. J. Campbell, Chem. Phys. Lett. 287, 143 1998 . 30 M. J. Shultz, C. Schnitzer, D. Simonelli, and S. Baldelli, Int. Rev. Phys. Chem. 19, 123 2000 . 31 C. Schnitzer, S. Baldelli, and M. Shultz, J. Phys. Chem. B 104, 585 2000 . 32 D. Zhang, J. Gutow, and K. Eisenthal, J. Phys. Chem. 98, 13729 1994 . 33 D. Zhang, J. Gutow, and K. B. Eisenthal, J. Chem. Soc., Faraday Trans. 92, 539 1996 . 34 K. R. Wilson, M. Cavalleri, B. S. Rude, R. D. Schaller, A. Nilsson, L. M. Pettersson, N. Goldman, T. Catalano, J. Bozek, and R. J. Saykally, J. Phys.: Condens. Matter 14, L221 2002 . 35 K. R. Wilson, R. D. Schaller, D. T. Co, R. J. Saykally, B. S. Rude, T. Catalano, and J. Bozek, J. Chem. Phys. 117, 7738 2002 . 36 H. Held, A. I. Lvovsky, X. Wei, and Y. R. Shen, Phys. Rev. B 66, 205110 2002 . 37 X. Wei, S.-C. Hong, A. I. Lvovsky, H. Held, and Y. R. Shen, J. Phys. Chem. B 104, 3349 2000 . 38 A. Morita, Chem. Phys. Lett. 398, 361 2004 . 39 M. Bonn, H. Ueba, and M. Wolf, J. Phys.: Condens. Matter 17, S201 2005 . 40 R. Kubo, in Fluctuation, Relaxation and Resonance in Magnetic Systems, edited by D. T. Haar Oliver and Boyd, Edinburgh and London, 1961 . 41 X. Wei and Y. R. Shen, Phys. Rev. Lett. 86, 4799 2001 . 42 R. DeVane, C. Ridley, T. Keyes, and B. Space, J. Chem. Phys. 119, 6073 2003 . 43 R. DeVane, C. Ridley, A. Perry, C. Neipert, T. Keyes, and B. Space, J. Chem. Phys. 121, 3688 2004 . 44 P. Moore, H. Ahlborn, and B. Space, in Liquid Dynamics Experiment, Simulation and Theory, edited by M. D. Fayer and J. T. Fourkas ACS Symposium Series, New York, 2002 . 45 H. Ahlborn, X. Ji, B. Space, and P. B. Moore, J. Chem. Phys. 112, 8083 2000 . 46 X. Ji, H. Ahlborn, B. Space, P. Moore, Y. Zhou, S. Constantine, and L. D. Ziegler, J. Chem. Phys. 112, 4186 2000 .

Downloaded 20 Mar 2012 to 130.102.158.22. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/about/rights_and_permissions

Potrebbero piacerti anche